International Journal of Molecular Sciences

Research Article
Biological Evaluation of 2-Aminobenzo[de]isoquinolin-1,3diones Against Herpes Simplex HSV-1 and 2
Rashad Al-Salahi 1* Ibrahim Alswaidan1, Essam Ezzeldin2 and Mohamed
Marzouk 1,3*
1
Department of Pharmaceutical Chemistry, 2Drug Bioavailability Lab.,
College of Pharmacy, King Saud University, P.O. Box 2457, Riyadh
11451, Saudi Arabia
3
Chemistry of Natural Products Group, Center of Excellence for Advanced
Sciences, National Research Center, Dokki, 12622, Cairo, Egypt
correspondence should be addressed to Rashad Al-salahi;[email protected]
and Mohamed Marzouk; [email protected]
As part of our search for new compounds having antiviral effects, the
prepared 2-aminonaphthalimide series was examined for its activity
against HSV-1 and HSV-2 viruses. This represents the first study of the
antiviral effects of this class of compounds. The new series of 2-amino1H-benzo[de]isoquinolin-1,3-diones was examined against herpes
simplex viruses HSV-1 and HSV-2 using a cytopathic effect inhibition
assay. In terms of effective concentration (EC50), furaldehyde (14),
thiophene aldehyde (15) and allyl isothiocyanide (16) derivatives showed
potent activity against HSV-1 (EC50= 19.6, 16.2 and 17.8 μg/mL), with
respect to acyclovir as a reference drug (EC50= 1.8 μg/mL). Moreover, 14
and 15 were found to exhibit valuable activity against HSV-2. Many of
the tested compounds demonstrated weak to moderate EC50 values
relative to their inactive parent compound (2-amino-1Hbenzo[de]isoquinolin-1,3-dione), while compounds 7, 9, 13, 14, 15, 16,
21 and 22 were the most set populated antiviral active compounds
throughout this study. The cytotoxicity (CC50), EC50, and the selectivity
index (SI) values were determined. Based on the potent anti-HSV
properties of the previous naphthalimide condensate products, further
exploration of this series of 2-amino-1H-benzo[de]isoquinolin-1,3-diones
is warranted.
1. Introduction
HSV-1 and HSV-2, as DNA viruses, belong to the alpha herpes virus subfamily,
which also includes the varicella zoster virus (VZV). They are common human
pathogens; between 60% and 95% of certain populations are infected with HSV-1,
and between 6% and 50% with herpes 2 (HSV-2) [1,2]. Herpes simplex viruses are
2
common pathogens that also cause herpes genitalis, herpes labialis, encephalitis and
keratitis. The infection caused by the two types is mainly transmitted by close
personal contact, and the virus establishes lifelong latent infection in the sensory
neurons, with recurrent lesions [3,4]. The frequency of HSV-seropositive males is
significantly higher in populations infected with the human immunodeficiency virus
(HIV). Moreover, sexually transmitted diseases like genital HSV increase the risk of
transmission and acquisition of HIV infection [5]; there is also a synergistic
relationship between genital herpes and HIV [6]. It was reported that HSVsuppressive therapy greatly reduced genital and plasma levels in patients co-infected
by HIV-1 RNA [7]. Hence, reducing the spread of genital herpes can greatly
decrease the risk of acquiring or transmitting HIV infection.
Imides of aromatic dicarboxylic acids like naphthalimides are important in the
construction of macromolecules as well as in supramolecular assembly. They are
useful fluoroprobes for various studies and also serve as precursors for the protection
of the amino group [8]. The 1H-benzo[de]isoquinolin-1,3-diones are also found to
play an important role as synthon for the construction of many bioactive compounds
such as antitumour and histone deacetylase inhibitors (HDAC) [9-16]. Moreover,
they have a high binding affinity towards the 5-HT1A receptor that is expressed in
CHO cells, as determined by fluorescence microscopy [17]. In our recent research,
the titled compounds were evaluated for their antimicrobial and cytotoxicity
activities, whereby some of them were found to possess significant effects [18,19]. In
spite of promising findings in the literature, most such compounds have been poorly
studied because of their complicated synthesis process [11]. In view of these facts,
we aimed to investigate a new prepared series of 2-amino-benzo[de]isoquinolin-1,3diones as antiviral agents against HSV-1 and HSV-2.
2. Materials and Methods
2.1. Mammalian cell line
Vero cells (derived from the kidney of a normal African green monkey) were
obtained from the American Type Culture Collection (ATCC). The viral strains used
were GHSV-UL46 for HSV-1 and the G strain for HSV-2. The vero cells were
propagated in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with
10% heat-inactivated foetal bovine serum (FBS), 1% L-glutamine, HEPES buffer
3
and 50 µg/mL gentamycin. All cells were maintained at 37 ºC in a humidified
atmosphere with 5% CO2 and were subcultured two times a week [19,20].
2.2. Evaluation of the antiviral activity using cytopathic effect inhibition assay
The antiviral screening was performed using a cytopathic effect inhibition assay
at the Regional Center for Mycology and Biotechnology RCMB, Al-Azhar
University, Cairo, Egypt [20]. This assay was selected to show specific inhibition of
a biologic function, that is, a cytopathic effect in susceptible mammalian cells [21].
In brief, monolayers of 10,000 vero cells adhering at the bottom of the wells in a 96well microtiter plate were incubated for 24 h at 37 ºC in a humidified incubator with
5% CO2. The plates were washed with fresh DMEM and challenged with 104 doses
of herpes simplex 1 or 2 virus, and then the cultures were simultaneously treated with
half-fold serial dilutions of the tested compound, starting from 500 µg/mL and going
down to about 2 µg/mL (500, 250, 125, 62.5, …, 1.95 µg/mL) in a fresh maintenance
medium; following this, they were incubated at 37 ºC for 48 h. An infection control,
as well as an untreated vero cell control was made in the absence of tested
compounds. Six wells were used for each concentration of the tested compound.
Every 24 h, an observation was made under the inverted microscope until the virus in
the control wells showed complete viral-induced cytopathic effects. Antiviral activity
was determined by the inhibition of the cytopathic effect compared to a control, that
is, the protection offered by the tested compound to the cells was scored [22]. Three
independent experiments were assessed, each containing four replicates per
treatment. Acyclovir, which is clinically used for the treatment of herpetic viral
disease, was used as positive control in this assay system [23].
After the incubation period, the media were aspirated, and then the cells were
stained with a 1% crystal violet solution for 30 min. Thereafter, all excess stain was
removed by rinsing the plates with tap water. The plates were allowed to dry, and
then glacial acetic acid (30%) was added to all wells and mixed thoroughly; the
absorbance of the plates was measured after gentle shaking on a microplate reader
(TECAN, Inc.), at 590 nm [20]. The viral inhibition rate was calculated as follows:
[(ODtv-ODcv)/(ODcd-ODcv)] × 100%, where ODtv, ODcv and ODcd indicate
the absorbance of the tested compounds with virus-infected cells, the absorbance of
the virus control and the absorbance of the cell control, respectively.
4
From these data, the dose that inhibited viral infection by 50% (EC50) was
estimated with respect to the virus control from the graphic plots, using STATA
modelling software. EC50, was determined directly from the curve obtained by
plotting the inhibition of the virus yield against the concentration of the samples. The
selectivity index (SI) was calculated from the ratio of CC50 to EC50 in order to
determine whether each compound had sufficient antiviral activity that exceeded its
level of toxicity [24]. This index is referred to as a therapeutic index, and it was also
used to determine whether a compound warranted further study. Compounds that had
an SI-value of 2 or more were considered to be active.
2.3. Cytotoxicity evaluation using viability assay
The vero cell lines in the cytotoxicity assay were seeded in 96-well plates at a
cell concentration of 1×104 cells per well in 100 µL of the growth medium [19,22].
Fresh medium containing different concentrations of the tested sample was added
after 24 h of seeding. Serial two-fold dilutions of the tested compound were added to
confluent cell monolayers dispensed into 96-well, flat-bottomed microtiter plates
(Falcon, NJ, USA) using a multichannel pipette. The microtiter plates were incubated
at 37 ºC in a humidified incubator with 5% CO2 for a period of 48 h. Three wells
were used for each concentration of the tested sample. Control cells were incubated
without test samples and with or without DMSO. The small percentage of DMSO
present in the wells (maximal 0.1%) was not found to affect the experiment.
After the end of the incubation period, the viable cell yield was determined by a
colorimetric method. In brief, the media were aspirated, and the crystal violet
solution (1%) was added to each well for at least 30 min. The stain was removed, and
the plates were rinsed using tap water until all excess stain was removed. Glacial
acetic acid (30%) was then added to all wells and mixed thoroughly. Absorbance of
the glacial acetic acid solution in each well was then measured at 590 nm using a
microplate reader (SunRise, TECAN, Inc, USA). The absorbance was proportional to
the number of surviving cells in the culture plate. All the results were corrected for
background absorbance detected in wells without added stain. Treated samples were
compared with the cell control in the absence of the tested compounds. All
experiments were carried out in triplicate. The cell cytotoxic effect of each tested
compound was calculated [23,25].
2.4. Data analysis
5
Statistics were done using a one-way ANOVA test [26], and the percentage cell
viability was calculated using Microsoft Excel®, as follows:
The % Cell viability = [(Mean Abscontrol Mean Abstest metabolite) / Mean Abscontrol] X
100, where Abs equals the absorbance at 590 nm. The 50% cell cytotoxic
concentration (CC50), the concentration required to kill or cause visible changes in
50% of intact mammalian cells, was estimated from graphic plots. The STATA
statistical analysis package was used for the dose response curve drawing, in order to
calculate CC50. Concerning antiviral evaluation, all experiments and data were
performed and analysed by Dr Mahmoud M. Elaasser (Regional Center for
Mycology and Biotechnology RCMB, Al-Azhar University, Cairo, Egypt).
3. Results and Discussion
In our previous research, we have described the synthetic routes and full
characterization of compounds 1-24, as illustrated in Scheme 1 and Table 1 [27].
Scheme 1. Main routes for synthesis of the target compounds 1-24.
O
O
a
N
N
N
NH2
R
O
O
c
1-15
b
O
O
S
N
N
N
H
O
16-18
R
NHR
O
19-24
a: aldehydes, DMF, refluxing for 5.5 hrs, b: isothiocyanides, DMF, refluxing for 6 hrs,
c: acid anhydrides, glacial acetic acid, refluxing for 6 hrs.
The present work demonstrates the evaluation of in vitro antiviral activity for the
new prepared 2-aminobenzo-[de]isoquinolin-1,3-diones against HSV-1 and HSV-2
by means of a cytopathic effect inhibition assay. From the obtained results, it can be
seen that the target molecules were found to possess weak to moderate anti-HSV
activity. The in vitro results of tested samples of both HSV-1 and HSV-2 were
recorded in Table 2, where target compounds 14, 15 and 16 showed remarkable and
significant activity against HSV-1, with EC50 values of 19.6, 16.2 and 17.8 μg/mL,
respectively, with respect to acyclovir (1.8 μg/mL) and their inactive parent
compound (2-amino-1H-benzo[de]isoquinolin-1,3-dione, Scheme 1). Furthermore,
6
14 and 15 exhibited good activity against HSV-2, giving EC50 values of 71.8 and
56.7 μg/mL, respectively, with respect to acyclovir (3.4 μg/mL). Compounds 7, 9, 21
and 22 displayed good activity against HSV-1 (EC50 = 113.08, 78.3, 102 and 74.6
μg/mL, respectively) in comparison with the previous active compounds.
Additionally, compound 9 demonstrated good effect against HSV-2, relative to 14
and 15, where it gave an EC50 value of 108 μg/mL.
Table 1. Synthesised 2-amino-benzo[de]isoquinolin-1,3-dione derivatives (1-24)
Cpd.
1
R
Cpd.
13
H3C CHCH-
2
R
N
CH-
14
CHO
15
CH-
3
CHS
O
4
16
CH 2-
CH-
O
17
5
CH-
CH-
6
18
O2N
O
7
CH-
HO
19
N
O
O
8
CH-
N
20
N
O
O
O
9
CH-
HO
21
N
O
O
O
10
CH-
O
22
N
O
Cl
11
CH-
Cl
23
O
O
O
N
O
O
OH
12
CHBr
24
O O
O
N N
N
O
O
O
7
Depending on the activity of 14 and 15, the products 3-6, 10, 17, 18 and 23 can
be considered as structures that are less active against HSV-1, while 4, 7, 10, 13, 16,
21 and 22 are less active against HSV-2. In terms of SI-values, all investigated
compounds can be sorted into two groups: inactive (SI < 2) and active (SI ≥ 2).
Accordingly, compounds 4, 5, 7, 9, 10, 13, 16, 17 and 21-23 can be considered as
active agents. Compounds 14 and 15 were found to be the most active compounds
against HSV-1, and similarly, 9, 14 and 15 could be accounted the only active
compounds against HSV-2 (Table 2). The present cytotoxicity evaluation of target
products 1-24 revealed a variety in activity, in which compounds 14, 15 and 16
yielded the highest effects against HSV-1, while 7, 9, 21 and 22 manifested a good
antiviral effect. On the other hand, 5, 6, 10, 17, 18 and 23 showed low activity
against HSV-1, while 4, 7, 10, 13, 16 and 21 showed low activity against HSV-2, in
comparison with compounds 14-16.
Regarding the elaborated results, biological studies showed that the type of
substituent produced from a condensation reaction with the amino group is a
controlling factor governing all of the observed pharmacological properties in the
parent structure. This is seen in the increment of antiviral activity in the sequence 15,
16 and 14 against HSV-1, whereas compound 15 showed higher activity than 14
against HSV-2. This could be attributed to the magnitude and conformation of the
hetero aldehyde and isothiocyanide substituent, which could have a substantial effect
on the activity and selectivity profiles of such compounds. Throughout this study, we
noticed that variation in the substituent on the benzyl ring demonstrated a remarkable
change in the activity profile. Regarding this fact, it was found that compounds 7 and
9 appeared as the most active among all aromatic aldehydes 3-7, 9 and 10 against
HSV-1. In addition, compound 9 seemed to be most highly active against HSV-1
from the group of compounds that also include 4, 7 and 10. Concerning the
isothiocyanide derivatives, compound 16 was more active than 17 and 18 against
HSV-1. Finally, the antiviral activity increased in the order of 22 > 21 > 23 against
HSV-1 and 21 > 22 against HSV-2 in the case of acid anhydride derivatives.
The cytopathic effect inhibition assay showed that some of the examined
compounds have good antiviral activity against HSV-1 and 2 in vitro at noncytotoxic concentrations with respect to their EC50 and SI, compared to acyclovir. In
order to understand the temporal aspects of the antiviral activity of target molecules,
it could be suggested that the mode of action of such compounds is not the
8
prevention of viral adsorption or penetration, but perhaps the interference in early
events of HSV replication, including immediate early (IE) transcriptional events. The
possible mode and mechanism of action of the tested molecules on early events of
the viral infection cycle could be attributed to the effect of compounds on
interference with the IE gene expression of HSV-1 and -2 [28,29]. The actual
explanation for these changes regarding structure-activity relationship will await the
elucidation of the mechanism(s) of action of the compounds.
Table 2. Antiviral activity of compounds 1-24 in terms of CC50, EC50 and SI against
HSV-1 and HSV-2.
Cpd Nr.
CC50
HSV-1
HSV-2
EC50
SI
EC50
SI
1
375 ±25
500
-
500
-
2
317 ±13
500
-
500
-
3
196 ±22
158 ± 1.8
1.2
500
-
4
486 ±18
189 ±3.2
2.6
312 ±9.6
1.6
5
452 ±9
176 ±4.6
2.6
500
-
6
316 ±11
174 ±1.8
1.8
500
-
7
412 ±24
113 ±4.6
3.6
220 ±3.8
1.9
8
289 ±13
500
-
500
-
9
236 ±8
78.3 ±1.5
3.0
108 ±4.1
2.2
10
380 ±16
149 ±5.3
2.6
416 ±8.2
0.9
11
316 ±24
500
-
500
-
12
320 ±18
500
-
500
-
13
458 ±27
117.2 ±3.1
3.9
308 ±7.1
1.5
14
321 ±23
19.6 ±0.9
16.4
71.8 ±1.8
4.5
15
180 ±8
16.2 ±1.1
11.1
56.7 ±2.6
3.2
16
96 ±12
17.8 ±1.4
5.4
292 ±4.8
0.3
17
518 ±34
236 ±4.9
2.2
500
-
18
482 ±14
319 ±3.7
1.5
500
-
19
220 ±16
500
-
500
-
20
220 ±16
500
-
500
-
21
490 ±21
102 ±5.8
4.8
374 ±3.7
1.3
22
560 ±46
74.6 ±3.4
7.5
428 ±10.2
1.3
23
412 ±18
176 ±2.4
2.3
500
-
24
386 ±17
500
-
500
-
Parent
320 ±24
500
-
500
-
Acyclovir
600 ±18
1.8 ±0.2
333.33
3.4 ±0.6
176.47
Cells treated with DMSO (0.1%) were used as a negative control, and its reading was
subtracted from the readings of tested compounds. Parent = 2-amino-1Hbenzo[de]isoquinolin-1,3-dione. Statistics were calculated using one-way ANOVA.
9
4. Conclusions
This study has revealed that compounds 14, 15 and 16 are active agents against
both herpes simplex viruses. These compounds could be useful as templates for
furthering development and design more potent antiviral agents.
Acknowledgments
The authors extend their appreciation to the Deanship of Scientific Research at
King Saud University for funding this work through research group No RGP-291.
Conflicts of Interest
The authors declare no conflict of interest.
Authors’ Contribution
Rashad A. Al-Salahi has made a substantial contribution to experimental design,
significant contribution to acquisition of data, manuscript preparation (writing and
revision) and approved the final form of manuscript. Ibrahim A. Alswaidan has
participated in reading and revision processes and approved the final form of
manuscript. Essam Ezzeldin has written and analyzed the statistical data and
participated in the revision process and approved the final form of manuscript.
Mohamed S. Marzouk has made a substantial contribution to experimental design,
significant contribution to acquisition of data, manuscript preparation (writing and
revision) and approved the final form of manuscript.
References
[1] J. S. Smith, and J. N. Robinson, “Age-specific prevalence of infection with
herpes simplex virus types 1 and 2: a global review,” The Journal of
Infectious. Diseases, vol. 186, no. 1, pp. S3-S28, 2002
[2] P. Akanitapichata, A. Wangmaneerata, P. Wilairatb, and K.F. Bastowc, “Antiherpes virus activity of Dunbari abella Prain,” Journal of Ethnopharmacology,
vol. 105, no. 1-2, pp. 64-68, 2006
[3] M. Fatahzadeh, and R. A. Schwartz, “Human herpes simplex virus infections:
epidemiology, pathogenesis, symptomatology, diagnosis, and management,”
10
Journal of the American Academy of Dermatology,
vol. 57, no. 5, pp. 737-
763, 2007
[4] F. M. Cowan, R. S. French, P. Mayaud, R. Gopal et al., “Seroepidemiological
study of herpes simplex virus types 1 and 2 in Brazil, Estonia, India, Morocco,
and Sri Lanka,” Sexually Transmitted. Infections, vol. 79, no. 4, pp. 286 -290,
2003
[5] S. Safrin, C. Crumpacker, P. Chatis, R. Davis et al., “A controlled trial
comparing foscarnet with vidarabine for acyclovir-resistant muco cutaneous
herpes simplex in the acquired immunodeficiency syndrome,”
The New
England Journal of Medicine, vol. 325, no. 8, pp. 551-555, 1991
[6] M. K. White, T. S. Gorrill, and K. Khalili, “Reciprocal transactivation between
HIV-1 and other human viruses,” Virology, vol. 352, no. 1, pp.1-13, 2006
[7] A. Wald, “Synergistic interactions between herpes simplex virus type-2 and
human immunodeficiency virus epidemics,” Herpes, vol.11, no. 3, pp. 70–76,
2004
[8] N. Barooah, C. Tamuly, and J. B. Baruah, “Synthesis, characterisation of few Nsubstituted 1,8-naphthalimide derivatives and their copper(II) complexes,”
Journal of Chemical Sciences, vol. 117, no. 2, pp.117–122, 2005
[9] Y. Xu, B. Qu, X. Qian, and Y. Li, “Five-member thio-heterocyclic fused
naphthalimides with aminoalkyl side chains: intercalation and photocleavage to
DNA,” Bioorganic Medicinal Chemistry Letters, vol. 15, no. 4, pp. 1139–1142,
2005
[10] A. Mukherjee, S. Hazra, S. Dutta, S. Muthiah et al., “Antitumor efficacy and
apoptotic activity of substituted chloroalkyl 1H-Benzo[de]Isoquinoline-1,3diones: A new class of potential antineoplastic agents,” Investigational New
Drugs, vol. 29, no. 3, pp. 434–442, 2011
[11] M. F. Brana, J. M. Castellano, M. Moran, M. J. Pérez de Vega et al., “ Bisnaphthalimides: a new class of antitumor agents,” Anti-Cancer Drug Design,
vol. 8, no. 4, pp. 257–268, 1993
11
[12] W. Aibin, L. L Jide, Q. Shaoxiong, and M. Ping, “ Derivatives of 5-nitro-1H
benzo[de]isoquinolin-1,3(2H)-dione: design, synthesis, and biological activity,”
Monatsh Chemie, vol. 141, pp. 95–99, 2010
[13] A. Mukherjee, S. Dutta, M. Shanmugavel, D. M. Mondhe et al., “6-Nitro-2-(3
hydroxypropyl)-1H- benz[de]isoquinoline-1,3-dione, a potent antitumor agent,
induces cell cycle arrest and apoptosis,” Journal of . Experimental & Clinical
Cancer Research, vol. 29, pp. 175-182. 2010
[14] G. N. Qazi, A. K. Saxena, S. Muthiah, D. M. Mondhe et al., “Preparation of
benzisoquinolinedione derivatives for us as antitumor agents,”
WO
2008084496 A1 20080717, 2008
[15] W. M. Cholody, T. Kosakowska-Cholody, and C. J. Michejda, “Preparation of
1,8-naphthalimido-linked
imidazo[4,5,1-de]acridones
as
bis-intercalating
antitumor agents,” WO 2001066545 A2 20010913, 2001
[16] A. Vaisburg, N. Bernstein, S. Frechette, M. Allan, et al., “(2-Aminophenyl)amides of omega-substituted alkanoic acids as new histone deacetylase
inhibitors,” Bioorganic Medicinal Chemistry Letters, vol. 14, no. 1, pp. 283–
287, 2004
[17] E., Lacivita, M. Leopoldo, A. C. Masotti, C. Inglese et al., “ Synthesis and
characterization of environment-sensitive fluorescent ligands for human 5HT1A receptors with 1-arylpiperazine structure,” Journal of
Medicinal
Chemistry, vol. 52, no. 23, pp. 7892–7896, 2009
[18] R. Al-Salahi, M. Marzouk, “Some 2-amino-benzo[de]isoquinolin-1,3-diones as
antimicrobial agents,” Asian Journal of Chemistry, vol. 26, no. 23, pp. 81638165, 2014
[19] R. Al-Salahi, I. Alswaidan, M. Marzouk, “Cytotoxicity Evaluation for a New
Set of 2-Aminobenzo[de]iso-quinolin-1,3-diones,” International Journal of
Molecular Sciences, vol. 15, pp. 22483-22491, 2014
[20] R. Al-Salahi, M. Marzouk, I. Alswaidan, and M. Al-Omar, “Antiviral activity
of 2-Phenoxy-4H-[1,2,4]triazolo[1,5-a]quinazoline derivatives,” Life Science
Journal, vol. 10, no. 4, pp. 2164–2169, 2013
12
[21] J. M. Hu, G. D. Hsiung, “Evaluation of new antiviral agents I: In vitro
prospectives,” Antiviral Research, vol. 11, no. 5-6, pp. 217-232, 1989
[22] P. Vijayan, C. Raghu, G. Ashok, S. .A. Dhanaraj, and B. Suresh, “Antiviral
activity of medicinal plants of Nilgiris,” Indian Journal of Medical Research,
vol. 120, no. 1, pp. 24-29, 2004
[23] D. J. Dargan, “Investigation of the anti-HSV activity of candidate antiviral
agents,” In: Methods in Molecular Medicine, Vol. 10, Herpes Simplex Virus
Protocols. Eds.: S.M. Brown, A.R. MacLean and N.J. Totowa: Humana Press
Inc, p. 387-405, 1998
[24] K. Zandi, M. A. Zadeh, K. Sartavi, and Z. Rastian, “Antiviral activity of Aloe
vera against herpes simplex virus type 2: An in vitro study,” African Journal of
Biotechnology Vol. 6, no. 15, pp. 1770-1773, 2007
[25] T. Mosmann, “Rapid colorimetric assay for cellular growth and survival:
Application
to
proliferation
and
cytotoxicity
assays,”
Jouranl
of
Immunological Methods, 65, no.1-2, pp. 55-63, 1983
[26] L. Castilla-Serna, and J. Cravioto, “Simply Statistic for Health Investigation,”
1st ed., Trillas: Mexico, D.F., 1999
[27]
R.
Al-Salahi,
and
M.
Marzouk,
aminobenzo[de]isoquinolin-1,3-dione
“Synthesis
derivatives,”
of
Asian
novel
2-
Journal
of
Chemistry, vol. 26, no. 15, pp. 2166-2172, 2014
[28] P. Bag, D. Ojha, H. Mukherjee, U. C. Halder et al., “An indole alkaloid from
a tribal folklore inhibits immediate early event in HSV-2 infected cells with
therapeutic efficacy in vaginally infected mice,” Pols One, vol. 8, no. 12, pp.1–
17, 2013
[29] Y. Liang, J. L. Vogel, A. Narayanan, H. Peng, and T. M. Kristie, “Inhibition of
the histone demethylase LSD1 blocks α-herpesvirus lytic replication and
reactivation from latency,” Nature Medicine, vol. 15, no. 11, pp. 1312-1317,
2009
.